International
Tables for
Crystallography
Volume D
Physical properties of crystals
Edited by A. Authier

International Tables for Crystallography (2006). Vol. D. ch. 3.3, pp. 412-414

Section 3.3.7.1. Growth twinning

Th. Hahna* and H. Klapperb

a Institut für Kristallographie, Rheinisch–Westfälische Technische Hochschule, D-52056 Aachen, Germany, and bMineralogisch-Petrologisches Institut, Universität Bonn, D-53113 Bonn, Germany
Correspondence e-mail:  hahn@xtal.rwth-aachen.de

3.3.7.1. Growth twinning

| top | pdf |

Growth twins can occur in nature (minerals), in technical processes or in the laboratory during growth from vapour, melt or solution. Two mechanisms of generation are possible for growth twins:

  • (i) formation during nucleation of the crystal;

  • (ii) formation during crystal growth.

3.3.7.1.1. Twinning by nucleation

| top | pdf |

In many cases, twins are formed during the first stages of spontaneous nucleation, possibly before the sub-critical nucleus reaches the critical size necessary for stable growth. This idea was originally proposed by Buerger (1945[link], p. 476) under the name supersaturation twins. There is strong evidence for twin formation during nucleation for penetration and sector twins, where all domains originate from one common well defined `point' in the centre of the twinned crystal, which marks the location of the spontaneous nucleus.

Typical examples are the penetration twins of iron borate FeBO3 (calcite structure), which are intergrowths of two rhombohedra, a reverse and an obverse one, and consist of 12 alternating twin domains belonging to two orientation states (see Example 3.3.6.5[link] and Fig. 3.3.6.4[link]). Experimental details are presented by Klapper (1987[link]) and Kotrbova et al. (1985[link]). Further examples are the penetration twins of the spinel law (Example 3.3.6.6[link] and Fig. 3.3.6.6[link]), the very interesting and complex [001] penetration twin of the monoclinic feldspar orthoclase (Fig. 3.3.7.1[link]) and the sector twins of ammonium lithium sulfate with three orientation states (Fig. 3.3.7.2[link]).

[Figure 3.3.7.1]

Figure 3.3.7.1 | top | pdf |

Orthoclase (monoclinic K-feldspar). Two views, (a) and (b), of Carlsbad penetration twins (twofold twin axis [001]).

[Figure 3.3.7.2]

Figure 3.3.7.2 | top | pdf |

Photographs of (001) plates ([\approx] 20 mm diameter, [\approx] 1 mm thick) of NH4LiSO4 between crossed polarizers, showing sector growth twins due to metric hexagonal pseudosymmetry of the orthorhombic lattice. (a) Nearly regular threefold sector twin (three orientation states, three twin components). (b) Irregular sector twin (three orientation states, but five twin components).

It should be emphasized that all iron borate crystals that are nucleated from flux or from vapour (chemical transport) exhibit penetration twinning. The occurrence of untwinned crystals has not been observed so far. Crystals of isostructural calcite and NaNO3, on the other hand, do not exhibit penetration twins at all. In contrast, for ammonium lithium sulfate, NH4LiSO4, both sector-twinned and untwinned crystals occur in the same batch. In this case, the frequency of twin formation increases with higher supersaturation of the aqueous solution.

The formation of contact twins (such as the dovetail twins of gypsum) during nucleation also occurs frequently. This origin must always be assumed if both partners of the final twin have roughly the same size or if all spontaneously nucleated crystals in one batch are twinned. For example, all crystals of monoclinic lithium hydrogen succinate precipitated from aqueous solution form dovetail twins without exception.

The process of twin formation during nucleation, as well as the occurrence of twins only for specific members of isostructural series (cf. Section 3.3.8.6[link]), are not yet clearly understood. A hypothesis advanced by Senechal (1980[link]) proposes that the nucleus first formed has a symmetry that is not compatible with the lattice of the (macroscopic) crystal. This symmetry may even be noncrystallographic. It is assumed that, after the nucleus has reached a critical size beyond which the translation symmetry becomes decisive, the nucleus collapses into a twinned crystal with domains of lower symmetry. This theory implies that for nucleation-twinned crystals, a metastable modification with a structure different from that of the stable macroscopic state may exist for very small dimensions. For this interesting theoretical model no experimental proof is yet available, but it appears rather reasonable; as a possible candidate of this kind of genesis, the rutile `eightling' in Example 3.3.6.9[link] may be considered.

Recently, the ideas on twin nucleation have been experimentally substantiated by HRTEM investigations of multiple twins. The formation of these twins in nanocrystalline f.c.c. and diamond-type cubic materials, such as Ge, Ag and Ni, is explained by the postulation of various kinds of noncrystallographic nuclei, which subsequently `collapse' into multiply twinned nanocrystals, e.g. fivefold twins of Ge; cf. Section 3.3.10.6.[link] An extensive review is provided by Hofmeister (1998[link]).

3.3.7.1.2. Twinning during crystal growth

| top | pdf |

  • (a) An alternative theory of twinning postulates the formation of a two-dimensional nucleus in twin position on a growth face of an existing macroscopic (previously untwinned) crystal. Such a mechanism was extensively described by Buerger (1945[link], pp. 472–475) and followed up by Menzer (1955[link]) and Holser (1960[link]). Obviously, this process is favoured by defects (inclusions, impurities) in the growth face. If the twin nucleus spreads out over the entire growth face, the twin boundary coincides with the growth face. This mechanism is generally assumed for the generation of large-area lamellar polysynthetic growth twins as observed for albite (Example 3.3.6.11[link] and Fig. 3.3.6.11[link]). For a critical discussion of the origin of irrational twin interfaces in rotation twins such as the pericline twins see Cahn (1954[link], p. 408). It should be noted that this mechanism is possible only for twin boundaries of very low energy, since the boundary energy of the large interface has to be supplied in one step, i.e. during spreading out of one growth layer in twin position. It is obvious that this kind of twin formation can only occur if the twin boundary coincides with a prominent growth face (F-face, rarely S-face, according to Hartman, 1956[link]).

  • (b) In the majority of growth twins, the twin boundary does not coincide with the growth face. This is the rule for merohedral twins, where the twin domains appear as `inserts' in the shape of pyramids or lamellae extending from the initiating defects (mostly inclusions) into the direction of growth of the face on which the twin has started. Examples are the pyramid-shaped Brazil-twin inserts of quartz (Frondel, 1962[link], Fig. 61 on p. 87) and the lamellar stripes of growth twins of KLiSO4 (Klapper et al., 1987[link], especially. Fig. 5). Similar pyramidal twin inserts are observed for Dauphiné growth twins in natural and synthetic quartz. These twin morphologies in quartz, however, are often considerably modified after growth by (partial) ferrobielastic switching of the domains, which is easily induced by stress at elevated temperatures [cf. Section 3.3.7.3[link](iii)[link]]. Illustrations of such Dauphiné twins are given by Frondel (1962[link], Fig. 49 on p. 78).

    The growth-twin inserts as described above appear improbable for non-merohedral twins because unfavourable high-energy boundaries would be involved. As a consequence, it must be concluded that non-merohedral twins with boundaries not coinciding with a (prominent) growth face (e.g. dovetail twins of gypsum) must form during the nucleation stage of the crystal [Section 3.3.7.1.1[link] above].

  • (c) Another model of twin formation has been suggested by Schaskolsky & Schubnikow (1933[link]). It is based on the idea that in the melt or solution the pre-existing small crystals make accidental contact with analogous faces [(hkl)_1] and [(hkl)_2] parallel, rotate and agglutinate in twin position, and continue to grow as a twin. This concept is also favoured by Buerger (1960b[link]). The model of Schaskolsky & Schubnikow is based on their interesting experiments with many ([\approx 1400]) K-alum crystals (up to 0.5 mm in size), which sediment in solution on horizontal octahedron (111) and cube (100) faces of large alum crystals (20–30 mm in size). A statistical analysis of the orientation distribution of the sedimented crystals reveals a significantly increased frequency of (111)/(111) parallel intergrowths, of regular (001)/(111) intergrowths and of (111) spinel twins. The authors interpret this result as a rotation of the small crystals around the contact-face normal after deposition on the large crystal. This initial contact plane (ICP) model of twin formation was critically discussed by Senechal (1980[link]) and considered as questionable, an opinion which is shared by the present authors.

  • (d) Finally, it is pointed out that twinning may drastically modify the regular growth morphology of (untwinned) crystals. A prominent example is the tabular shape of (111)-twinned cubic crystals with the large face parallel to the (111) contact plane. This is due to the increased lateral growth rate of the faces meeting in re-entrant edges (re-entrant corner effect; Hartman, 1956[link]; Ming & Sunagawa, 1988[link]). The (111) tabular shape of twinned cubic crystals plays an important role for photographic materials such as silver bromide, AgBr (Buerger, 1960b[link]; Bögels et al., 1997[link], 1998[link]). A more extreme habit modification is exhibited by the [\langle110\rangle] growth needles of cubic AgBr, which contain two [\{111\}] twin planes intersecting along [\langle110\rangle] (Bögels et al., 1999[link]).

    The phenomenon of habit modification by twinning has been developed further by Senechal (1976[link], 1980[link]), who presents an alternative model of the genesis of penetration twins (cf. Section 3.3.7.1.1[link] above): initial cubic (111) contact twins consisting of two octahedra change their habit during growth so as to form two interpenetrating cubes of the spinel law. As a further example, chabasite is cited.

  • (e) During melt growth of the important cubic semiconductors with the diamond structure (Si, Ge) and sphalerite (zinc sulfide) structure (e.g. indium phosphide, InP), twins of the spinel law [twin mirror plane (111) or twofold twin axis [111], cf. Section 3.3.10.3.3[link]] are frequently formed. Whereas this twinning is relatively rare and can easily be avoided for Si and Ge, it is a persistent problem for the III–V and II–VI compound semiconductors, especially for InP and CdTe crystals, which have a particularly low {111} stacking-fault energy (Gottschalk et al., 1978[link]). For Czochralsky growth, these twins are usually nucleated at `edge facets' forming at the surface of the `shoulder' (or `cone region') where the growing crystal widens from the seed rod to its final diameter. Once nucleated, they proceed during further growth as bulk twins or, more frequently, as twin lamellae with sharp [\{111\}] contact planes. For a [111] pulling direction, the three equivalent {111} twin planes with inclination of 19.5° against the pull axis [111] are usually activated, whereas the perpendicular (111) twin plane does not or only rarely occurs (Bonner, 1981[link]; Tohno & Katsui, 1986[link]). These twins can be avoided by optimizing the growth conditions, in particular by the choice of a proper cone angle, which is the most crucial parameter. A mechanism of the {111} twin formation of III–V compound semiconductors was suggested by Hurle (1995[link]) and experimentally confirmed for InP, using synchroton-radiation topography combined with chemical etching and Normarski microscopy, by Chung et al. (1998[link]) and Dudley et al. (1998[link]). A comprehensive X-ray topographic study of (111) twinning in indium phosphide crystals, grown by the liquid-encapsulated Czochralski technique, and its interaction with dislocations is presented by Tohno & Katsui (1986[link]).

  • (f) It should be noted lastly that `annealing twins' (which are an important subject in metallurgy) are not treated in this section, because they are considered to be part of bicrystallography. These twins are formed during recrystallization and grain growth in annealed polycrystalline materials (cf. Cahn, 1954[link], pp. 399–401).

References

First citation Bonner, W. A. (1981). InP synthesis and LEC growth of twin-free crystals. J. Cryst. Growth, 54, 21–31.Google Scholar
First citation Buerger, M. J. (1945). The genesis of twin crystals. Am. Mineral. 30, 469–482.Google Scholar
First citation Buerger, M. J. (1960b). Introductory remarks. Twinning with special regard to coherence. In Symposium on twinning. Cursillos y Conferencias, Fasc. VII, pp. 3 and 5–7. Madrid: CSIC.Google Scholar
First citation Bögels, G., Buijnsters, J. G., Verhaegen, S. A. C., Meekes, H., Bennema, P. & Bollen, D. (1999). Morphology and growth mechanism of multiply twinned AgBr and AgCl needle crystals. J. Cryst. Growth, 203, 554–563.Google Scholar
First citation Bögels, G., Meekes, H., Bennema, P. & Bollen, D. (1998). The role of {100} side faces for lateral growth of tabular silver bromide crystals. J. Cryst. Growth, 191, 446–454.Google Scholar
First citation Bögels, G., Pot, T. M., Meekes, H., Bennema, P. & Bollen, D. (1997). Side-face structure and growth mechanism of tabular silver bromide crystals. Acta Cryst. A53, 84–94.Google Scholar
First citation Cahn, R. W. (1954). Twinned crystals. Adv. Phys. 3, 202–445.Google Scholar
First citation Chung, H., Dudley, M., Larson, D. J., Hurle, D. T. J., Bliss, D. F. & Prasad, V. (1998). The mechanism of growth-twin formation in zincblende crystals: insights from a study of magnetic-liquid encapsulated Czochralski grown InP single crystals. J. Cryst. Growth, 187, 9–17.Google Scholar
First citation Dudley, M., Raghothamachar, B., Guo, Y., Huang, X. R., Chung, H., Hurle, D. T. J. & Bliss, D. F. (1998). The influence of polarity on twinning in zincblende structure crystals: new insights from a study of magnetic liquid-encapsulated Czochralski-grown InP crystals. J. Cryst. Growth. 192, 1–10.Google Scholar
First citation Frondel, C. (1962). The system of mineralogy, 7th edition, Vol. III. Silica minerals, especially pp. 75–99. New York: Wiley.Google Scholar
First citation Gottschalk, H., Patzer, G. & Alexander, H. (1978). Stacking fault energy and ionicity of cubic III–V compounds. Phys. Status Solidi A, 45, 207–217.Google Scholar
First citation Hartman, P. (1956). On the morphology of growth twins. Z. Kristallogr. 107, 225–237.Google Scholar
First citation Hofmeister, H. (1998). Forty years study of fivefold twinned structures in small particles and thin films. Cryst. Res. Technol. 33, 3–25, especially Section 4.Google Scholar
First citation Holser, W. T. (1960). Relation of pseudosymmetry to structure in twinning. In Symposium on twinning. Cursillos y Conferencias, Fasc. VII, pp. 19–30. Madrid: CSIC.Google Scholar
First citation Hurle, D. T. J. (1995). A mechanism for twin formation during Czochralski and encapsulated vertical Bridgman growth of III–V compound semiconductors. J. Cryst. Growth, 147, 239–250.Google Scholar
First citation Klapper, H. (1987). X-ray topography of twinned crystals. In Progress in crystal growth and characterization, Vol. 14, edited by P. Krishna. pp. 367–401. Oxford: Pergamon.Google Scholar
First citation Klapper, H., Hahn, Th. & Chung, S. J. (1987). Optical, pyroelectric and X-ray topographic studies of twin domains and twin boundaries in KLiSO4. Acta Cryst. B43, 147–159.Google Scholar
First citation Kotrbova, M., Kadeckova, S., Novak, J., Bradler, J., Smirnov, G. V. & Shvydko, Yu. V. (1985). Growth and perfection of flux-grown FeBO3 and 57FeBO3 crystals. J. Cryst. Growth, 71, 607–614.Google Scholar
First citation Menzer, G. (1955). Über Kristallzwillingsgesetze. Z. Kristallogr. 106, 193–198.Google Scholar
First citation Ming, N. B. & Sunagawa, I. (1988). Twin lamellae as possible self-perpetuating step sources. J. Cryst. Growth, 87, 13–17.Google Scholar
First citation Schaskolsky, M. & Schubnikow, A. (1933). Über die künstliche Herstellung gesetzmässiger Kristallverwachsungen des Kalialauns. Z. Kristallogr. 85, 1–16.Google Scholar
First citation Senechal, M. (1976). The mechanism of formation of certain growth twins of the penetration type. Neues Jahrb. Mineral. Monatsh. pp. 518–525.Google Scholar
First citation Senechal, M. (1980). The genesis of growth twins. Sov. Phys. Crystallogr. 25, 520–524.Google Scholar
First citation Tohno, S. & Katsui, A. (1986). X-ray topographic study of twinning in InP crystals grown by the liquid-encapsulated Czochralski technique. J. Cryst. Growth, 74, 362–374.Google Scholar








































to end of page
to top of page